• Nie Znaleziono Wyników

Seismic interferometry applied to local fracture seismicity recorded at Planchón-Peteroa Volcanic Complex, Argentina-Chile

N/A
N/A
Protected

Academic year: 2021

Share "Seismic interferometry applied to local fracture seismicity recorded at Planchón-Peteroa Volcanic Complex, Argentina-Chile"

Copied!
36
0
0

Pełen tekst

(1)

Delft University of Technology

Seismic interferometry applied to local fracture seismicity recorded at Planchón-Peteroa

Volcanic Complex, Argentina-Chile

Casas, J. A.; Draganov, D.; Badi, G. A.; Manassero, M. C.; Olivera Craig, V. H.; Franco Marín, L.; Gómez, M.; Ruigrok, E. DOI 10.1016/j.jsames.2019.03.012 Publication date 2019 Document Version

Accepted author manuscript Published in

Journal of South American Earth Sciences

Citation (APA)

Casas, J. A., Draganov, D., Badi, G. A., Manassero, M. C., Olivera Craig, V. H., Franco Marín, L., Gómez, M., & Ruigrok, E. (2019). Seismic interferometry applied to local fracture seismicity recorded at Planchón-Peteroa Volcanic Complex, Argentina-Chile. Journal of South American Earth Sciences, 92, 134-144. https://doi.org/10.1016/j.jsames.2019.03.012

Important note

To cite this publication, please use the final published version (if applicable). Please check the document version above.

Copyright

Other than for strictly personal use, it is not permitted to download, forward or distribute the text or part of it, without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license such as Creative Commons. Takedown policy

Please contact us and provide details if you believe this document breaches copyrights. We will remove access to the work immediately and investigate your claim.

(2)

Seismic interferometry applied to regional and teleseismic

events recorded at Planch´

on-Peteroa Volcanic Complex,

Argentina-Chile

Casas, Jos´e Augustoa, Draganov, Deyanb, Badi, Gabriela Alejandrac, Franco, Luisd

a

Facultad de Ciencias Astron´omicas y Geof´ısicas, Universidad Nacional de La Plata, CONICET, Argentina

bDepartment of Geoscience and Engineering, Delft University of Technology, The

Netherlands

cFacultad de Ciencias Astron´omicas y Geof´ısicas, Universidad Nacional de La Plata,

Argentina

dObservatorio Volcanol´ogico de los Andes del Sur (OVDAS-SERNAGEOMIN), Chile

Abstract

The Planch´on-Peteroa Volcanic Complex (PPVC) is located in the Cen-tral Andes, Argentina-Chile. Even though this active volcanic system is considered one of the most dangerous volcano in the region, with more than twenty modest (VEI < 4) Holocene eruptions, knowledge of its subsurface structures, internal processes, dynamics, and their relation, is still limited.

Seismic interferometry (SI) is a high-resolution technique based on anal-yses of the interference of the propagated seismic energy at one or many stations. SI can be used to characterize the subsurface properties of a target area. In particular, previous SI studies performed in the area of the PPVC describe specific ranges of depth; therefore, more information is required for a thorough description of the subsurface features in the area and for a better understanding of the PPVC dynamics.

We apply SI based on autocorrelations of selected regional and tele-seismic events to image the subsurface structures below stations located in

© 2018 Manuscript version made available under CC-BY-NC-ND 4.0 license https://creativecommons.org/licenses/by-nc-nd/4.0/

(3)

Argentina and Chile during 2012. The selection of the events is performed according to their location, magnitude, angle of incidence of P-wave seismic energy, and signal to noise ratio in the records. For each station, we extract time windows and we process them using three ranges of frequency, which are sensitive to different ranges of depths.

This work describes depths and zones previously not analyzed in the area. The results not only complement the available geological, geochemical, and geophysical information, but present new information for depths between 5 and ∼400 km depth, increasing the general knowledge of the subsurface features in the PPVC. Finally, we also propose a model for the first 45 km of the subsurface (i.e., down to the Moho), which indicates the crustal structure and the likely distribution of magma bodies in depth.

Keywords:

Planch´on-Peteroa Volcanic Complex, Seismic Interferometry, Regional and teleseismic events, Magma storage in depth

1. Introduction 1

The Planch´on-Peteroa Volcanic Complex -PPVC- (35.223◦S, 70.568◦W; 2

see location inFigure 1) is located in the Andes at the international border 3

between Argentina and Chile. The PPVC is composed of three main volcanic 4

edifices, i.e., the Azufre, the Planch´on, and the Peteroa, out of which the 5

latter is the current active volcano. The PPVC presents overlapped calderas 6

originated from the destruction of several volcanic structures during past 7

explosive events (Tormey,1989). Through analyses of its historical activity 8

and products, this volcanic system is ranked as the most hazardous volcano 9

in Argentina (Elissondo and Far´ıas,2016) and the eighth most risky volcano 10

(4)

in Chile (Technical sheet, Observatorio Volcanol´ogico de los Andes del Sur, 11

OVDAS-SERNAGEOMIN, Chile). 12

The knowledge of the PPVC has been developed by the contribution 13

from several disciplines, i.e., geology (Tormey, 1989; Haller et al., 1994; 14

Naranjo et al.,1999;Tapia Silva,2010;Haller and Risso,2011), geochemistry 15

(Benavente, 2010; Tassi et al., 2016; Benavente et al., 2016), meteorology 16

(Guzm´an et al.,2013), ash analysis (Ramires et al.,2013), seismology (Casas 17

et al.,2014;Manassero et al.,2014;Olivera Craig,2017;Casas et al.,2018; 18

Casas et al.), gravimetry (Tassara et al., 2006), and risk analysis (Haller 19

and Coscarella). These studies contribute to the knowledge of the eruptive 20

history and the current subsurface conditions of this volcanic system. Nev-21

ertheless, the dynamics the PPVC and their relation with the subsurface 22

structures are still poorly understood, increasing the local risk (Elissondo 23

and Far´ıas,2016). 24

A description of the subsurface structures (i.e., depth, associated dimen-25

sions, density contrasts, etc.) is essential for developing accurate knowl-26

edge of the dynamics of any volcanic system. In particular, knowledge of 27

subsurface discontinuities provides constraints for tomographic studies, for 28

magma-ascent modeling, among others, contributing to a better inference 29

of the subsurface conditions, and, therefore, leading to more reliable analy-30

ses of likely future volcanic scenarios. Based on structural-geology analyses, 31

Tapia Silva (2010) describes the subsurface geological units located in the 32

very first 10 km of the subsurface in the area of the PPVC, and present 33

their distribution in depth. Even though no local studies have been applied 34

for describing the crustal structure in the PPVC, Far´ıas et al. (2010) and 35

Giambiagi et al. (2012) provide a crustal structure as a function of depth 36

and the distance from the trench in the Central Andes. For the depth of 37

(5)

the subducting slab below the PPVC, they estimate four zones delimited in 38

depth at ∼12 (the intracrustal discontinuity), ∼27, and 45 km depth -the 39

crust-mantle discontinuity (the Moho). The Moho is estimated at ∼45 km 40

depth (Tassara et al.,2006); the intra-lithosphere discontinuity (top of litho-41

spheric low-velocity zone), at ∼70 km depth (Karato,2012); and the top of 42

the subducting slab, at ∼120 km depth (Tassara et al.,2006). Nevertheless, 43

more scientific evidence is required to increase the information about the 44

known subsurface structures, leading to a more accurate characterization of 45

their properties, as well as to describe the subsurface features previously not 46

analyzed. These goals motivate local studies, as the one presented in this 47

article. 48

Claerbout (1968) has constituted a frame over which the theory of seis-49

mic interferometry developed. This passive seismic method -from here on, 50

Seismic Interferometry by Autocorrelations (SIbyA)- suggests that the au-51

tocorrelation of a plane-wave transmission response propagating in a hori-52

zontally layered medium, recorded at the surface, allows the retrieval of the 53

reflection response of a virtual source co-located to the recording station. 54

SIbyA has shown to be a robust method; it has been applied to different 55

type of seismic data, in several areas and at different scales. For example, 56

SIbyA was applied to global- and teleseismic phases to imaging the crustal 57

subsurface at regional scales (Ruigrok and Wapenaar,2012;Nishitsuji et al., 58

2016), to P-wave of microseismic events to imaging the shallow volcanic sub-59

surface (Kim et al.,2017), and to ambient-noise seismic data at several scales 60

(Draganov et al.,2007; Gorbatov et al.,2013;Boullenger et al.,2014;Oren 61

and Nowack,2017). The robustness of SIbyA has motivated its application 62

to local (Casas et al.), regional, and teleseismic seismic data recorded in the 63

area of the PPVC. 64

(6)

Nishitsuji et al. (2016) apply SIbyA to global seismic phases recorded 65

in the eastern flank of the Peteroa volcano during 2012. They confirm the 66

location of the Moho at ∼45 km depth, and propose a deformation feature 67

in the subducting slab in the form of detachment, shearing, necking, or any 68

combination of them. 69

Casas et al. apply SIbyA to local seismic events to image the subsurface 70

below the stations located in the Argentine and Chilean sides of the PPVC 71

during 2012. They confirm the geological structure described for the first 72

4 km of the subsurface (Tapia Silva, 2010), provide information about re-73

gions of higher heterogeneity caused by faulting and complex geochemical 74

processes, and support the presence of a magma body emplaced at ∼4 km 75

depth (previously suggested byBenavente (2010)). 76

We apply SIbyA to regional and teleseismic events selected according to 77

their location, magnitude, angles of incidence of the P-wave seismic energy 78

at each station, and the signal to noise ratio in the records. The results 79

for three different frequency ranges allow the description of the subsurface 80

structures between ∼5 and 400 km depth, and the inference of the crustal 81

structure and the location of magma bodies down to the Moho. 82

2. Data 83

The present application uses seismic data recorded by stations deployed 84

in Argentina and Chile during 2012 (see station distribution in Figure 1). 85

The temporal deployment of seismic instruments in an area of interest is 86

a widely used tool for reaching several goals, e.g., perform first analyses of 87

the propagated wavefield and the subsurface conditions, increase the num-88

ber of the recording stations, extend the analyzed area, and improve the 89

(7)

accuracy of the results. The MalARRgue project (Ruigrok et al.,2012) was 90

designed by institutions from The Netherlands (Delft University of Technol-91

ogy -TUDelft), Argentina (Comisi´on Nacional de Energ´ıa At´omica CNEA), 92

and The United States (Boise State University -BSU). Its goal is imaging 93

and monitoring the subsurface of the Malarge region (Mendoza, Argentina), 94

an area of high scientific interest due to peculiar volcanic and tectonic pro-95

cesses (Stern, 2004). The MarlARRgue project consisted in a temporal 96

deployment (from January 2012 to January 2013) of 38 stations, out of 97

which six were deployed along the eastern flank of the PPVC (from here 98

on, the PV array). The PV array was equipped with short-period (2 Hz) 99

three-component (Sercel L-22) sensors. 100

Another source of data is provided by three broad-band stations of the 101

Observatorio Volcanol´ogico de los Andes del Sur (OVDAS-SERNAGEOMIN, 102

Chile), which are located ∼6 km northwards. These stations (from here on, 103

OVDAS array) were active during 2012, through the same period as the PV 104

array. 105

3. Application and results 106

SIbyA is described by the reciprocity theorem of correlation type ( Wape-107

naar,2003,2004). Based on this theorem for transient sources (Wapenaar 108

and Fokkema,2006), and using autocorrelation in the time domain, we ob-109 tain: 110 X sources T (xA, −t) ∗ si(−t) ∗ T (xA, t) ∗ si(t) ⊗ s(−t) ∗ s(t)i ≈ −R(xA, −t) + δ(t) − R(xA, t) , (1)

(8)

which states that the reflection response R(xA, t) can be retrieved at

111

the station A located (at xA) at the surface through the autocorrelation

112

of a recorded transmitted wavefield T (xA, t). The operator ∗ indicates

113

convolution, ⊗ means deconvolution, and δ is the Dirac’s delta. The fac-114

tor s(−t) ∗ s(t)i corresponds to the autocorrelated source time function 115

(ASTF), which allows the deconvolution of each source time function si(t).

116

Even thoughEquation 1requires sources over the whole stationary phase 117

area (i.e., the Fresnel Zone), seismic events present a non-uniform spatial 118

distribution. Therefore, performing a selection of the seismic sources to be 119

used is essential for a proper application of SIbyA. In order the transmission 120

response of the propagated seismic energy to be accurately estimated by 121

the vertical component of the records, we select only seismic events with 122

P-wave seismic energy arriving (sub) vertically to the station at the surface. 123

The retrieved reflection response (from here on, Rv(xA, t)) is related to a

124

seismic source co-located to the station at the surface, radiating P-wave 125

energy vertically downwards. 126

A seismic source in the subsurface release energy that propagates to-127

wards the surface in which it is reflected. This seismic energy is reflected, 128

refracted, and diffracted at the subsurface structures and heterogeneities (or 129

the surface), part of which arrives to the recording station at the surface. 130

Seismograms are then composed of direct waves followed by these reverber-131

ated waves. SIbyA removes the times previous to the direct arrival, and 132

attenuates the incoherent noise, providing seismic evidence of the location 133

of the subsurface structures. Figure 2 pictures the application of SIbyA in 134

an idealized horizontally layered 2-D medium, given a plane wavefield orig-135

inated by a seismic source located exactly below the station. The obtained 136

reflection response can be used to know the depth of the reflectors located 137

(9)

in the subsurface below the station. 138

In the real Earth, nor the wave fronts are plane at local and regional 139

scales nor the subsurface is horizontally layered in volcanic zones. In highly 140

heterogeneous zones (as, for example, the area of the PPVC;Manassero et al. 141

(2014)), the location of a seismic source exactly below the station is not an 142

imperative condition for an accurate retrieval of the subsurface reflection 143

response Rv(xA, t) (Fan and Snieder, 2009), i.e., the vertical component

144

of the records is still an accurate estimation of the transmission response. 145

Therefore, sources with small P-wave angles of incidence are selected. 146

3.1. Pre-processing 147

This section aims to get the input data and prepare it for the proper 148

application of theEquation 1. Using the reference seismic catalogs (IRIS and 149

USGS), we select events occurred during the recording period (i.e., January 150

2012 until January 2013) and which are characterized by a sufficiently great 151

magnitude so that signal to noise ratio is high in the records of each station. 152

Due to likely variations of the local seismic wavefield in space and time, we 153

judge the signal to noise ratio of each event at each of the stations. 154

For the selection of seismic events, we use the software JWEED (Java 155

version of Windows Extracted from Event Data) developed by IRIS. Based 156

on restrictions in the origin time, the location, and the magnitude, we pre-157

select events (seeFigure 3). According to their epicentral distance, we clas-158

sify them in two groups. One group is composed of events with epicentral 159

distances between 30◦ and 120◦, and magnitudes greater than Mw. 6; each 160

event of this group guarantees a sufficiently small P-wave ray parameter 161

(< 0.08 s/km) so that seismic energy arrives (sub)vertically at the station, 162

i.e., with incident angles <∼ 25◦ (Kennett et al., 1995). The second group 163

(10)

is composed of events with epicentral distances lower than 30◦ and magni-164

tudes greater than Mw. 5. These events present a wide range of possible 165

P-wave angles of incidence. Therefore, we perform an examination analysis 166

on this second group in order to select only those events with at least one 167

P-wave phase smaller than the adopted threshold (i.e., 0.08 s/km). The ray 168

parameters estimated by the regional velocity model ak135 (Kennett et al., 169

1995) are appropriate for this analysis, as smaller angles of incidence of the 170

P-wave energy are expected in the real Earth, provided its higher hetero-171

geneity (Fan and Snieder,2009). Once the seismic events are selected, there 172

is no need to keep the distinction between the groups, i.e., they are equally 173

significant. 174

The origin time of the selected events is used to extract the seismic 175

waveforms from the records of the PV and OVDAS stations. A first estimate 176

of the P- and S-wave arrival times for each event is calculated using the 177

regional velocity model ak135. These times are then employed to manually 178

pick accurate P- and S-wave arrival times, as well as to get the frequency 179

range of a sufficiently high signal-to-noise ratio. We request a good (> 180

0.8) signal to noise ratio for the events to be processed, in order to avoid 181

non-interested high amplitudes. 182

Provided the origin time of the selected events, obtained the accurate 183

arrival times, and examined the (sub)vertical incidence of the P-wave en-184

ergy and high signal-to-noise ratio of the records, we extract the vertical-185

component records of the selected events at each of the used stations. 186

3.2. Processing 187

The vertical-component records of seismic events with P-wave energy 188

arriving vertically at a station represent an accurate estimate of the P-wave 189

(11)

transmission response of such propagated wavefield (provided the disconti-190

nuities are not excessively inclined;Nishitsuji et al. (2016)). 191

Out of the frequency range of processing previously selected for each 192

event according to its signal to noise ratio in the results, we use those fre-193

quencies greater than 0.3 Hz, a threshold defined by the instrumental char-194

acteristics of the PV-array stations (Nishitsuji et al., 2014). Furthermore, 195

we only use those frequencies shared through the events, i.e., [0.3 2.1] Hz. 196

In order to perform a better interpretation of the results through depth, we 197

segmented this frequency range in two sub-ranges, i.e., [0.3 0.8] Hz and [0.8 198

2.1] Hz. The separating frequency (0.8 Hz) is selected after a trial and error 199

approach, based on the observed coherency in the results for all the stations 200

in advanced stages of the processing. 201

In order to avoid the rise of non-physical arrivals caused by cross-terms 202

in the correlations, we extract the times between the first P-wave arrival 203

and the first S-phase arrival. As an example,Figure 4shows the processing 204

windows for the station PV04, for the complete range of frequencies (i.e., 205

[0.8 2.1] Hz). 206

The higher value of the selected frequency range (i.e., 2.1 Hz) restricts 207

the resolution of the results for particular depths. Therefore, out of the 208

(previously tested) vertically arriving seismic events, we make a third group 209

composed of those with epicentral distances smaller than 20◦. These events 210

are characterized by a sufficiently high signal-to-noise ratio up to 3.2 Hz. 211

As this group aims to provide information about shallower subsurface struc-212

tures, we select a minimum frequency of 1 Hz. Therefore, we apply the 213

same processing workflow to the three selected frequency ranges, i.e., [0.3 214

0.8] Hz, [0.8 2.1] Hz, and [1 3.2] Hz. As the same importance is assigned to 215

the events of each of the three groups, we normalize the processing windows 216

(12)

according to their vertical flux of seismic energy. 217

As suggested by Equation 1, we estimate and deconvolve the ASTF 218

from each of the autocorrelated time windows. The ASTF of each event is 219

estimated by the main lobe and the secondary monotonous-decreasing lobes, 220

as shown inFigure 5 for the station AD2 and the frequency range [0.3 0.8] 221

Hz. 222

Figure 6presents the autocorrelation of the time windows for the station 223

PV01 and the frequency range [0.3 0.8] Hz, in which each trace is decon-224

volved by its previously estimated ASTF. This figure shows the dominance 225

of the main lobe in the autocorrelated deconvolved traces. These features 226

close to 0 s are mainly non-physical amplitudes remaining from the decon-227

volution. Therefore, these amplitudes are removed through windowing. 228

SIbyA is based on the autocorrelation of time windows extracted from 229

the records of selected seismic events. Note that the autocorrelation of 230

a extracted time window could arise non-physical arrivals at times equal 231

to the time interval between two P-wave arrivals, reducing the quality of 232

the results. However, these time intervals are a function of the epicentral 233

distance of the events. The seismic events used in this application present 234

a wide range of epicentral distances, so that the non-physical arrivals are 235

located at different times in the autocorrelations, leading to a destructive 236

interference of their energy during stacking (seeFigure 7). 237

The last step in the application of Equation 1is stacking the resulting 238

seismic traces for each station, which enhances the energy from the sta-239

tionary phase area. Figure 8 shows the pre-stack panel (deconvolved and 240

windowed traces) and the stacked traces for stations AD2 and PV04, for the 241

three selected frequency ranges of processing. 242

(13)

4. Interpretation and discussion 243

Aiming to compare the seismic results with the known features of the 244

subsurface, we transform the time vector of the results to depth through 245

construction and utilization of a velocity model. This model is composed of 246

velocities provided by the regional model ak135 for depths greater than 60 247

km, and a modified version of the one obtained by Bohm et al. (2002) for 248

shallower depths (see used velocity model inFigure 9). 249

Figure 10 and Figure 11 show the stacked traces for the PV and OV-250

DAS arrays, respectively, for each processing frequency range. These figures 251

also show the interpreted subsurface features for each of the stations. As a 252

complex impedance contrast through depth is expected for the area of the 253

PPVC, we only seek for the dominant amplitudes on the obtained reflection 254

responses, which are potentially related to the main subsurface discontinu-255

ities. The lower frequency range (i.e., [0.3 0.8] Hz) leads to describe the 256

subsurface between ∼40 and 400 km depth. The results for the other two 257

frequency ranges (i.e., [0.8 2.1] Hz, and [1 3.2] Hz) allow to interpret the 258

subsurface features for depths between 5 and ∼45 km. The minimum depth 259

limit is set by the non-physical amplitudes removed from close to 0 s after 260

deconvolution. The maximum depth limit is set by the coherency in the 261

results for all the frequency ranges and all the used stations. 262

The interpretation of the results for the smallest frequency range ([0.3 263

0.8] Hz) is performed through contrast of the seismic results and the expected 264

location of the known subsurface features based on the geodynamic scenario 265

and the available geological information for the area of the PPVC (Ferr´an 266

and Mart´ınez,1962;Tassara et al.,2006;Benavente,2010;Tapia Silva,2010; 267

Karato,2012). 268

(14)

The results for the PV array (see Figure 10a) show six dominant ampli-269

tudes (i.e., local maximum on the absolute values of the waveforms), which 270

we classify as potential subsurface discontinuities. The close location of the 271

identified features in the seismic results and the known subsurface features 272

lead to the interpretation of the Mohorovicic discontinuity at ∼45 km depth, 273

the intra-lithospheric discontinuity at 65 km, the top of the subducting slab 274

between 110 and 120 km, the bottom of the subducting slab between 140 275

and ∼160 km, the lithosphere-asthenosphere boundary between 230 and 255 276

km, and the top of the asthenospheric low-velocity zone between ∼330 and 277

∼360 km depth. 278

The OVDAS array (see Figure 11a) is an array located ∼6 km to the 279

north of the PV array, composed of half the stations of the latter, and 280

with greater longitudinal extension. The results for the OVDAS array al-281

low to interpret the Mohorovicic discontinuity at ∼45 km depth, the intra-282

lithospheric discontinuity between 70 and 90 km, the top of the subducting 283

slab between 115 and 130 km, the bottom of the subducting slab between 284

∼165 and ∼185 km, the lithosphere-asthenosphere boundary at ∼250 km, 285

and the top of the asthenospheric low-velocity zone between ∼310 and ∼350 286

km depth. 287

Based on the seismic velocity values for the depths of interpretation and 288

the frequency range of processing, the resolution of the seismic results is 5 289

km (Widess, 1973). This value leads to interpret that the results for the 290

OVDAS array do not differ substantially from the results of the PV array, 291

what is expected provided the small geological variation in ∼6 km along 292

the north-south direction for the used processing wavelengths. The best 293

correlation in depth is observed for the Mohorovicic discontinuity (43-48 km 294

depth), the lithosphere-asthenosphere boundary (∼245 km), and the top 295

(15)

of the asthenospheric low-velocity zone (∼340 km). A small difference in 296

depth is observed for the intra-lithospheric discontinuity and the top of the 297

subducting slab; even though greater depths are observed in the results of 298

the OVDAS stations, these differences would not be significant based on 299

the vertical resolution of the results. A greater difference is observed for the 300

bottom of the subducting slab, i.e., ∼15 km greater for the OVDAS stations. 301

Although a dominant positive arrival is expected at the depth of the 302

Moho, a dominant negative amplitude is retrieved in the results for most 303

of the stations. Based on the retrieved waveforms, we interpret the pres-304

ence of a complex area at ∼40-55 km depth, causing a perturbation of the 305

amplitudes retrieved for these depths, in particular for those related to the 306

Moho. 307

Even though dipping structures in the subsurface restrict the reflection 308

energy arrived at the surface, we clearly recognize the depth of the top and 309

bottom of the subducting slab. Therefore, two hypotheses arise. One hy-310

pothesis suggests a stair-like subduction, according to which the top and the 311

bottom of the oceanic slab present horizontal (or gently inclined) regions; 312

the different depths estimated in the results of the PV and the OVDAS ar-313

rays for the bottom of the subducting slab could be caused by a local change 314

of the thickness of the subducting lithosphere. Nevertheless, this hypothesis 315

would not explain the lack of seismicity at the longitude of the stations and 316

depths of analysis (US Geological Survey;Nishitsuji et al.(2016)). A second 317

hypothesis (Nishitsuji et al.,2016) proposes a slab deformation in the form 318

of detachment, shearing, necking, or any combination. Then, a differential 319

deformation between the latitudes of the PV and OVDAS arrays would ex-320

plain the estimated depths for the bottom of the subducting slab. Finally, 321

more information is required to elucidate the proper interpretation. 322

(16)

For the two higher ranges of frequencies (i.e., [0.8 2.1] Hz and [1 3.2] Hz) 323

(seeFigure 10b,Figure 10c,Figure 11b, andFigure 11c), the interpretation 324

is also based on the identification of the dominant amplitudes in the results, 325

and the depths for which the arrived reflected energy is particularly smaller, 326

a feature probably caused by the emplacement of a sufficiently great volume 327

of magma as to be manifested in the seismic results. 328

The results for the PV array and the frequency range [0.8 2.1] Hz (see 329

Figure 10b) indicate five clear dominant arrivals in most of the stations, out 330

of which four are between ∼10 and ∼30 km depth and another one at ∼40 km 331

depth. Additionally, we identify an apparent lack of dominant amplitudes 332

for depths between ∼30 and ∼40 km (indicated with an arrow inFigure 10b). 333

The features identified for [0.8 2.1] Hz are supported by the results for the 334

frequency range [1 3.2] Hz (Figure 10c), which improve the depth of the 335

inferred subsurface discontinuities. In addition, these results manifest an 336

apparent low-amplitude region at ∼25 km depth for the western stations of 337

the array. The results for this frequency range also show a dominant arrival 338

at ∼6 km depth. 339

The results for the OVDAS stations agree with the interpretation per-340

formed for the PV array, for the two analyzed frequency ranges. Therefore, 341

we identify local-maximum amplitudes, as well as apparent small-amplitude 342

zones, at roughly the same depths for the two arrays and for the two higher-343

frequency ranges, even though the effect of attenuation increases for the 344

highest frequencies (around 3 Hz in this application) (Sch¨on,2015). Then, 345

these results allow the interpretation of the subsurface structures between 5 346

and ∼45 km depth (the Moho). 347

Based on the average depth of the reflectors interpreted in the seismic re-348

sults, the available scientific information about the subsurface in the PPVC, 349

(17)

the proposed structure of the crust for the Central Andes (Far´ıas et al., 350

2010;Giambiagi et al.,2012), and the physics of magma storage in the crust 351

Jackson et al. (2018), we propose a model for the distribution of magma 352

reservoirs in depth in relation to the main subsurface structures in the crust 353

(see Figure 12). 354

Through comparison of the average depth of the interpreted reflectors 355

below the stations and the proposed structure of the crust (Far´ıas et al., 356

2010;Giambiagi et al.,2012), we associate the interpreted reflectors at ∼12, 357

∼18, and ∼32 km depth as the intra-crustal discontinuity (rigid-ductile dis-358

continuity in the upper crust), the discontinuity between the upper and lower 359

crust, and the rigid-ductile discontinuity in the lower crust, respectively (see 360

Figure 12). 361

Jackson et al.(2018) models the formation, storage, and chemical differ-362

entiation of magma in the Earth’s crust. According to the physics of magma 363

storage, the melt fraction is not homogeneously distributed through depth. 364

A great percentage of melt is located in the very upper part of a reservoir, 365

a low percentage is located through most of the reservoir, while a solid area 366

is present in the lower part. The seismic results are most probably evidence 367

of the solid lower section of the reservoir (Jackson et al., 2018). Therefore, 368

we interpret a region in depth as characterized by a magma emplacement 369

in case two conditions are satisfied: 1. the presence of an area of smaller 370

amplitudes in the seismic results, and 2. it is located above any of the inter-371

preted subsurface reflectors. This circumstance is satisfied for two regions, 372

i.e., a shallower zone located above the rigid-ductile discontinuity in the 373

lower crust (i.e., ∼32 km depth); and a deeper one at ∼35 km depth, above 374

a reflector located at ∼40 km. 375

Even though no amplitude information is available for depths lower than 376

(18)

5 km depth (which are removed after deconvolution), a subsurface model for 377

the area (Benavente,2010) proposes a magma emplacement at ∼4 km depth. 378

We identify a reflector at ∼6 km depth, which motivates the incorporation 379

of such magma emplacement in our model. 380

Furthermore, two regions (indicated with a question mark in Figure 12) 381

satisfy only one of the imposed conditions, therefore, their interpretation as 382

regions of magma storage is subjected to extra information. These regions 383

are located above the reflectors interpreted at ∼22 depth and the Moho, 384

for which no apparent smaller amplitudes are observed, probably due to 385

its close location to another feature of the subsurface (upper-lower crust 386

discontinuity and the Moho, respectively), or the resolution of the seismic 387

results are not sufficiently great to recognize a region of limited vertical 388

extension of magma. 389

Our results support the information obtained for the subsurface in the 390

area (Yuan et al., 2006; Ward et al., 2013; Gonz´alez-Vidal et al., 2018) 391

which indicate (although with a limited resolution) low-velocity zones for 392

approximately the same range of depths. They are also consistent with 393

the conceptual model proposed for the area (Benavente, 2010) for depths 394

between 5 and 15 km depth, for which great volumes of magma storage are 395

not expected. 396

Finally, more research (e.g., local seismic velocity -or attenuation- tomog-397

raphy studies) is required to accurately identify the location and dimensions 398

of the regions of magma emplacement. 399

(19)

5. Conclusions 400

Even though the Planch´on-Peteroa Volcanic Complex (PPVC) is one of 401

the most hazardous volcanic systems in the Central Andes, knowledge of 402

its internal processes, structures, dynamics, and their relation are still not 403

satisfactorily understood. 404

We apply seismic interferometry by autocorrelations to regional and tele-405

seismic data recorded by nine stations deployed in the area of the PPVC (six 406

in Argentina and three in Chile) during 2012. The events are selected accord-407

ing their location, magnitude, angle of incidence of the P-wave energy, the 408

signal to noise ratio on the results, and the related useful frequency range. 409

In order to perform an appropriate description of the subsurface structures 410

below the stations, we use three frequency ranges ([0.3 0.8] Hz, [0.8 2.1] Hz, 411

and [1 3.2] Hz) which are sensitive to different range of frequencies. 412

The smallest frequency range ([0.3 0.8] Hz) is used to infer the tectonic 413

features, i.e., the Moho (at 43-48 km depth), the intra-lithospheric discon-414

tinuity (∼70 km), the top and bottom of the subducting slab (∼120 and 415

∼150-165 km), the lithosphere-asthenosphere boundary (∼250 km), and the 416

top of the asthenospheric low-velocity zone (∼340 km). The results support 417

the hypothesis of deformation in the form of detachment, searing, and/or 418

necking for the longitude of the used stations. Our results also suggest a 419

higher depth (∼15 km) for the bottom of the subducting slab at the north 420

of the PPVC, likely caused by differential deformation along the latitude 421

direction. 422

Based on the results for the two higher-frequency ranges ([0.8 2.1] Hz 423

and [1 3.2] Hz) and previous geological, geochemical, and geophysical infor-424

mation, we propose a model which describes the structure of the crust and 425

(20)

the subsurface regions storaging magma bodies down to the Moho. Three 426

regions of sufficiently great volume of magma emplaced at ∼4 km, ∼28 km, 427

and ∼35 km depth, respectively are indicated. 428

The present work provides valuable information about the subsurface 429

conditions of an active volcanic system -the CVPP. We expect the obtained 430

knowledge to be employed in future research aiming to better understand 431

the dynamics of the CVPP. 432

References 433

Benavente, O., 2010. Actividad Hidrotermal asociada a los Complejos 434

Volc´anicos Planch´on-Peteroa y Descabezado Grande-Quizapu-Cerro Azul, 435

36S y 37oS, Zona Volc´anica Sur, Chile. Universidad de Chile . 436

Benavente, O., Tassi, F., Reich, M., Aguilera, F., Capecchiacci, F., 437

Guti´errez, F., Vaselli, O., Rizzo, A., 2016. Chemical and isotopic fea-438

tures of cold and thermal fluids discharged in the Southern Volcanic Zone 439

between 32.5S and 36S: Insights into the physical and chemical processes 440

controlling fluid geochemistry in geothermal systems of Central Chile. 441

Chemical Geology . 442

Bohm, M., L¨uth, S., Echtler, H., Asch, G., Bataille, K., Bruhn, C., Riet-443

brock, A., Wigger, P., 2002. The Southern Andes between 36 and 40S 444

latitude: Seismicity and average seismic velocities. Tectonophysics . 445

Boullenger, B., Verdel, A., Paap, B., Thorbecke, J., Draganov, D., 2014. 446

Studying CO 2 storage with ambient-noise seismic interferometry: A com-447

bined numerical feasibility study and field-data example for Ketzin, Ger-448

many. Geophysics 80, Q1–Q13. 449

(21)

Casas, J.A., Badi, G., Manassero, M., Gomez, P., Draganov, D., Ruzzante, 450

J., 2014. Characterization of Seismo-volcanic Activity in Peteroa Volcano, 451

Central Andes Argentina-Chile. Earth Sciences Research Journal 18, 335– 452

336. 453

Casas, J.A., Draganov, D., Badi, G., Manassero, M.C., Olivera Craig, V., 454

Franco, L., G´omez, M., Ruigrok, E., . Seismic interferometry applied to 455

local fracture seismicity recorded at Planch´on-Peteroa Volcanic Complex, 456

Argentina-Chile. Manuscript under revision . 457

Casas, J.A., Mikesell, T.D., Draganov, D., Lepore, S., Badi, G.A., Franco, 458

L., G´omez, M., 2018. Shallow S-Wave Velocity Structure from Ambient 459

Seismic Noise at Planch´on-Peteroa Volcanic Complex, Argentina-Chile. 460

Bulletin of the Seismological Society of America 108, 2183–2198. 461

Claerbout, J.F., 1968. Synthesis of a layered medium from its acoustic 462

transmission response. GEOPHYSICS . 463

Draganov, D.S., Wapenaar, K., Mulder, W., Singer, J., Verdel, A., 2007. Re-464

trieval of reflections from seismic background-noise measurements. Geo-465

physical Research Letters 34. 466

Elissondo, M., Far´ıas, C., 2016. Volcanic Risk assessment in Argentina, in: 467

Cities on Volcanoes IX, Puerto Varas, Chile. 468

Fan, Y., Snieder, R., 2009. Required source distribution for interferometry 469

of waves and diffusive fields. Geophysical Journal International 179, 1232– 470

1244. 471

Far´ıas, M., Comte, D., Charrier, R., Martinod, J., David, C., Tassara, A., 472

Tapia, F., Fock, A., 2010. Crustalscale structural architecture in central 473

(22)

Chile based on seismicity and surface geology: Implications for Andean 474

mountain building. Tectonics 29. 475

Ferr´an, O.L.G., Mart´ınez, M.V., 1962. Reconocimiento geol´ogico de la 476

Cordillera de los Andes entre los paralelos 35 y 38 sur, in: Anales de 477

la Facultad de Ciencias F´ısicas y Matem´aticas, pp. ´ag–19. 478

Giambiagi, L., Mescua, J., Bechis, F., Tassara, A., Hoke, G., 2012. Thrust 479

belts of the southern Central Andes: Along-strike variations in shortening, 480

topography, crustal geometry, and denudation. Bulletin of the Geological 481

Society of America . 482

Gonz´alez-Vidal, D., Obermann, A., Tassara, A., Bataille, K., Lupi, M., 483

2018. Crustal model of the Southern Central Andes derived from ambient 484

seismic noise Rayleigh-wave tomography. Tectonophysics . 485

Gorbatov, A., Saygin, E., Kennett, B.L., 2013. Crustal properties from 486

seismic station autocorrelograms. Geophysical Journal International . 487

Guzm´an, C., Hucailuk, C., Tamasi, M., Mart´ınez Bogado, M., Torres, 488

D., 2013. Anomal´ıas Encontradas en los Par´ametros Registrados en la 489

Estaci´on de Medici´on de la Terma del Volc´an Peteroa, in: Actas de ICES 490

IX, pp. 186–194. 491

Haller, M.J., Coscarella, M., . An´alisis probabil´ıstico del riesgo de erupci´on 492

del volc´an Peteroa mediante la aplicaci´on de mezcla de distribuciones 493

exponenciales. Nat. Hazards Earth Syst. Sci 9, 425–431. 494

Haller, M.J., Ostera, H.A., Pesce, A.H., Gardini, M., Folguera, A., 1994. 495

Vulcanoestratigraf´ıa reciente y eruptividad del volc´an Peteroa, in: Con-496

greso Geol´ogico Chileno, pp. 319–323. 497

(23)

Haller, M.J., Risso, C., 2011. La erupci´on del volc´an peteroa (3515’s, 7018’o) 498

del 4 de septiembre de 2010. Revista de la Asociacion Geologica Argentina 499

. 500

Jackson, M.D., Blundy, J., Sparks, R.S., 2018. Chemical differentiation, 501

cold storage and remobilization of magma in the Earth’s crust. 502

Karato, S.I., 2012. On the origin of the asthenosphere. Earth and Planetary 503

Science LettersarXiv:1104.0048v2. 504

Kennett, B.L.N., Engdahl, E.R., Buland, R., 1995. Constraints on seismic 505

velocities in the Earth from traveltimes. Geophysical Journal International 506

. 507

Kim, D., Brown, L.D., ´Arnason, K., ´Agustsson, K., Blanck, H., 2017. 508

Magma reflection imaging in Krafla, Iceland, using microearthquake 509

sources. Journal of Geophysical Research: Solid Earth . 510

Manassero, M., Badi, G., Casas, J.A., Gomez, M., Draganov, D., Ruzzante, 511

J., 2014. Seismic attenuation around Peteroa Volcano, Argentina. Earth 512

Sciences Research Journal 18, 341–342. 513

Naranjo, J.A., Haller, M.J., Ostera, H.A., Pesce, A.H., Sruoga, P., 1999. 514

Geologia y peligros del Complejo Volc´anico Planch´on-Peteroa, Andes del 515

Sur (3515’S), Regi´on del Maule, Chile-Provincia de Mendoza, Argentina. 516

Servicio Nacional de Geologia y Mineria. 517

Nishitsuji, Y., Ruigrok, E., Gomez, M., Draganov, D., 2014. Global-phase 518

H/V spectral ratio for delineating the basin in the Malargue Region, Ar-519

gentina. Seismological Research Letters 85, 1004–1011. 520

(24)

Nishitsuji, Y., Ruigrok, E., Gomez, M., Wapenaar, K., Draganov, D., 2016. 521

Reflection imaging of aseismic zones of the Nazca slab by global-phase 522

seismic interferometry. Interpretation 4, SJ1–SJ16. 523

Olivera Craig, V., 2017. Relocation of fracture seismicity in Planch´ on-524

Peteroa Volcanic Complex through optimization of the arrival-times iden-525

tification and joint location techniques. Ph.D. thesis. 526

Oren, C., Nowack, R.L., 2017. Seismic body-wave interferometry using noise 527

autocorrelations for crustal structure. Geophysical Journal International 528

. 529

Ramires, A., Elissonde, A., Trombotto Liaudat, D., 2013. Posibles escenarios 530

de riesgo frente a la ca´ıda de cenizas volc´anicas, en el modelo ganadero 531

de la cuenca alta y media del Rio Grande, Malarg¨ue, Mendoza, in: Actas 532

de IX, pp. 304–320. 533

Ruigrok, E., Draganov, D., Gomez, M., Ruzzante, J., Torres, D., Pumarega, 534

I.L., Barbero, N., Ramires, A., Ganan, A.R.C., van Wijk, K., 2012. 535

Malarg¨ue seismic array: Design and deployment of the temporary array. 536

The European Physical Journal Plus 127, 126. 537

Ruigrok, E., Wapenaar, K., 2012. Global-phase seismic interferometry un-538

veils P-wave reflectivity below the Himalayas and Tibet. Geophysical 539

Research Letters 39. 540

Sch¨on, J.H., 2015. Physical properties of rocks: Fundamentals and principles 541

of petrophysics. volume 65. Elsevier. 542

Stern, C.R., 2004. Active Andean volcanism: its geologic and tectonic set-543

ting. Revista geol´ogica de Chile . 544

(25)

Tapia Silva, F.F., 2010. An´alisis estructural del sector occidental de la faja 545

plegada y corrida de Malarg¨ue en el curso superior del r´ıo Colorado de 546

Lontu´e (35 18’y 35o23’s), Regi´on del Maule, Chile. Universidad de Chile 547

. 548

Tassara, A., G¨otze, H.J., Schmidt, S., Hackney, R., 2006. Three-dimensional 549

density model of the Nazca plate and the Andean continental margin. 550

Journal of Geophysical Research: Solid Earth . 551

Tassi, F., Aguilera, F., Benavente, O., Paonita, A., Chiodini, G., Caliro, 552

S., Agusto, M., Gutierrez, F., Capaccioni, B., Vaselli, O., Caselli, A., 553

Saltori, O., 2016. Geochemistry of fluid discharges from Peteroa volcano 554

(Argentina-Chile) in 2010-2015: Insights into compositional changes re-555

lated to the fluid source region(s). Chemical Geology . 556

Tormey, D., 1989. Geologic history of the active Azufre-Planchon-Peteroa 557

volcanic center (3515’S, Southern Andes) with implications for the devel-558

opment of compositional gaps. Asoc. Gel. Arg. Rev , 420–430. 559

Wapenaar, K., 2003. Synthesis of an inhomogeneous medium from its acous-560

tic transmission response. GEOPHYSICS . 561

Wapenaar, K., 2004. Retrieving the elastodynamic Green’s function of an 562

arbitrary inhomogeneous medium by cross correlation. Physical Review 563

Letters 93. 564

Wapenaar, K., Fokkema, J., 2006. Green’s function representations for seis-565

mic interferometry. GEOPHYSICS . 566

Ward, K.M., Porter, R.C., Zandt, G., Beck, S.L., Wagner, L.S., Minaya, E., 567

(26)

Tavera, H., 2013. Ambient noise tomography across the Central Andes. 568

Geophysical Journal International . 569

Widess, M.B., 1973. How thin is a thin bed? GEOPHYSICS . 570

Yuan, X., Asch, G., Bataille, K., Bock, G., Bohm, M., Echtler, H., Kind, 571

R., Oncken, O., W¨olbern, I., 2006. Deep seismic images of the Southern 572

Andes. Geological Society of America Special Papers . 573

6. Figures 574

(27)

Figure 1: Distribution of the seismic stations used in the present application in relation to the main edifices of the Planch´on-Peteroa Volcanic Complex (PPVC).

(28)

Figure 2: Seismic interferometry by autocorrelation applied to vertically arriving energy in a horizontally layered medium. tj represents the two-way travel time between the

station at the surface and the reflector j in the subsurface. The autocorrelation allows the retrieval of a seismogram composed of reflected energy released by a virtual source co-located at the position of the station.

Figure 3: Location of seismic events pre-selected for the application of SIbyA in the area of the PPVC. A triangle indicates the location of the PPVC. Stars show the location of events with epicentral distances less than 30◦and magnitudes Mw > 5. Circles indicate events with epicentral distances greater than 30◦and less than 120◦, and magnitudes Mw > 6.

(29)

Figure 4: Processing time windows (P-wave codas) for each of the events selected for PV04 station in the complete range of frequencies, i.e., [0.3 2.1] Hz. Each window is normalized according to its vertical energy flux. Vertical axis indicates propagation time. Each window is composed of a pre-event time (20 s) and the times between the first P-and S-wave arrival times.

Figure 5: Autocorrelated source time functions (ASTFs) estimated for the station AD2 for the frequency range [0.3 0.8] Hz. A shaded area shows the ASTFs in the autocorrelation panel (for graphical purposes, we only show the first 15 s).

(30)

Figure 6: Autocorrelated time windows for the station PV01 in the frequency range [0.3 0.8] Hz. The vertical axis indicates two-way travel time. Each seismic trace is deconvolved by its previously estimated source time function.

Figure 7: Cartoon illustrating the attenuation of non-physical arrivals originated in the correlation of a time window with several P-wave arrivals. Stacking seismic traces from events with different epicentral distances enhances features located in phase, so that non-physical arrivals due to several P-wave arrivals are attenuated. Without loss of generality, this figure shows the effect of stacking using time windows of events with different epi-central distances, each of them composed of two P-wave phases. Ti is the time window

of the event i, which contains two P-wave arrivals separated in δti. Operator ∗∗ means

(31)

(a) AD2 [0.3 0.8] Hz

(b) AD2 [0.8 2.1] Hz

(32)

(d) PV04 [0.3 0.8] Hz

(e) PV04 [0.8 2.1] Hz

(f) PV04 [1 3.2] Hz

Figure 8: Pre-stacking panels and stacked seismic trace for the stations AD2 (a, b, c) and PV04 (d, e, f), for the frequency ranges [0.3 0.8] Hz (a, d), [0.8 2.1] Hz (b, e), and [1 3.2]

(33)

Figure 9: Velocity model used to perform the time-to-depth transformation of the seismic results.

(34)

(a)

(b) (c)

Figure 10: Interpretation of the results at the stations of the PV array for the three frequency ranges: (a) [0.3 0.8] Hz, (b) [0.8 2.1] Hz, y (c) [1 3.2] Hz. Filled rectangle areas show the local maximum amplitudes, i.e., the interpreted subsurface discontinuities below each station. Rectangles with dashed line borders indicate a higher uncertainty at the identification of a discontinuity. Discontinuities interpreted only in (c) are marked with a small circle in the left bottom corner of each rectangle. Figure 10calso shows the interpreted discontinuities at depths close to those interpreted in (b). Lm represent the

(35)

(a)

(b) (c)

Figure 11: Interpretation of the results at the stations of the OVDAS array for the three frequency ranges: (a) [0.3 0.8] Hz, (b) [0.8 2.1] Hz, y (c) [1 3.2] Hz. Filled rectangle areas show the local maximum amplitudes, i.e., the interpreted subsurface discontinuities below each station. Rectangles with dashed line borders indicate a higher uncertainty at the identification of a discontinuity. Discontinuities interpreted only in (c) are marked with a small circle in the left bottom corner of each rectangle. Figure 11calso shows the interpreted discontinuities at depths close to those interpreted in (b). Lm represent the

minimum depth level for the Moho (interpreted in (a)). Arrows indicate zones of likely emplacement of magma. The dashed arrow represents an uncertainty of interpretation

(36)

Figure 12: Proposed model of magma emplacement in relation to the structure of the crust down to the Moho in the area of the PPVC. Inverted triangles indicate the longitude of the stations. Thick horizontal lines below the stations show the average depth of the reflectors interpreted in the seismic results. Dashed lines are the interpreted discontinuities (based onFar´ıas et al. (2010) andGiambiagi et al.(2012)) between the different regions of the crust. Arrows show the inferred direction of magma movement. Areas with a question mark inside indicate zones of higher ambiguity in the interpretation.

Cytaty

Powiązane dokumenty

Although more scattered than in the publication cluster map ( figure 2 B), figure 4 shows a similar increasing trend in publications related to patient safety risk factors, as

Bogaty we wrażenia pierwszy dzień zjazdu zakończył się przy ognisku, które zapłonęło nad brzegiem Jeziora Lednickiego przy siedzibie dyrekcji Muzeum Pierwszych Piastów

It is unclear whether global network topology drives function, as suggested by some researchers, or whether it is simply a byproduct of evolution or even an artefact of

In that article, a publisher/subscriber based model is proposed, where three entities are recognized: an IoT application provider (e.g., a wireless sensor access network operator,

Analizując skład pierwiastkowy odpadu po pirolizie opon stwierdzono, że wysoka zawartość węgla (ok. 80% mas.) pozwoli na uzyskanie z badanego materiału sorbentu

W związku z tak dużą ilością danych i obiektów, funkcjami bezpieczeństwa czy sterowania, bardzo ważne stają się zagadnienia związane ze wspomi- nanym już bezpieczeństwem

Diesen Umstand kann man dadurch erklären, dass alle drei oben genannten Karten (vielleicht ist hier die Beatuskarte aus Saint-Sever eine Ausnahme) nach 1054 entstanden, also nach

– Oprawa: oryginalna, XVI-wieczna, organiczna, szyta na cztery podwójne zwięzy sznurkowe, okładziny drewniane obleczone brązową skórą, skóra wytarta ze spękaniami lica,